Toward the Synthesis of Spirastrellolide A: Construction of a

A stereocontrolled synthesis of the fully elaborated C26−C40 tricyclic [5,6,6]-bis-spiroacetal of spirastrellolide A containing the C28 chlorine sub...
0 downloads 0 Views 139KB Size
ORGANIC LETTERS

Toward the Synthesis of Spirastrellolide A: Construction of a Tetracyclic C26−C40 Subunit Containing the DEF-Bis-Spiroacetal

2005 Vol. 7, No. 19 4121-4124

Ian Paterson,* Edward A. Anderson, Stephen M. Dalby, and Olivier Loiseleur UniVersity Chemical Laboratory, UniVersity of Cambridge, Lensfield Road, Cambridge CB2 1EW, UK [email protected] Received June 16, 2005

ABSTRACT

A stereocontrolled synthesis of the fully elaborated C26−C40 tricyclic [5,6,6]-bis-spiroacetal of spirastrellolide A containing the C28 chlorine substituent is reported, exploiting asymmetric Sharpless dihydroxylation and boron-mediated allylation methodology.

Spirastrellolide A is a novel cytotoxic polyketide isolated by Andersen and co-workers from the Caribbean sponge Spirastrella coccinea in 2003.1 Biological testing has shown spirastrellolide A to be a potent (IC50 ) 1 nM) and selective inhibitor of protein phosphatase 2A (PP2A).1b As such, it may exhibit a similar mode of action to other Ser/Thr phosphatase inhibitors, including fostriecin, okadaic acid, and the calyculins, which induce premature cell mitosis.2 Spirastrellolide A represents a potential candidate for the development of antitumor therapeutic agents and, given the central regulatory role of PP2A in the cell, may find further applications as a lead in the treatment of neurological and metabolic disorders. Extensive NMR spectroscopic analysis of the methyl ester derivative of spirastrellolide led to a partial structural assignment and characterization of a 38-membered macrolactone core.1a Further structural analysis and derivatization revealed several inconsistencies in this original assignment and resulted in a second report from the Andersen group containing a revision of the structure of spirastrellolide A as 1 (Figure 1).1b (1) (a) Williams, D. E.; Roberge, M.; Van Soest, R.; Andersen, R. J. J. Am Chem. Soc. 2003, 125, 5296. (b) Williams, D. E.; Lapawa, M.; Feng, X.; Tarling, T.; Roberge, M.; Andersen, R. J. Org. Lett. 2004, 6, 2607. 10.1021/ol051403c CCC: $30.25 Published on Web 08/17/2005

© 2005 American Chemical Society

Figure 1.

Spirastrellolide contains 21 stereocenters, of which 20 are embedded within the 38-membered macrolactone. This ring

Scheme 1. Retrosynthetic Analysis of Spirastrellolide A

itself contains a cis-fused tetrahydropyran (A ring), a [6,6]spiroacetal (BC rings), and an intriguing chlorinated [5,6,6]bis-spiroacetal (DEF rings). Significantly, both spiroacetal motifs appear to benefit from stabilization by a double anomeric effect, with all substituents equatorially disposed. The remaining C46 stereocenter appears as part of a nine-carbon side chain, also featuring a (Z,E)-1,4-diene, appended to C38. While the atomic connectivity of spirastrellolide and its methyl ester derivative 2 has been determined, several stereochemical uncertainties remain. In addition to the remote, unassigned C46 stereocenter, the absolute configuration of the natural product remains undetermined. More significantly from a synthetic viewpoint,3 the relatiVe stereochemistry between the three stereoclusters (C3-C7, C9-C24, and C27-C38) within the macrolide core is also unknown. The initial aim of any synthetic effort directed toward spirastrellolide must be the elucidation of the stereochemical relationship between these subunits. A flexible, modular strategy was envisaged in which each region of known relative stereochemistry could be independently constructed. Subsequent union of these subunits, followed by detailed NMR comparisons with spirastrellolide, may then enable the determination of the stereochemical interconnection between these regions. Toward this end, we report herein an asymmetric synthesis of the C26-C40 segment 3 (Scheme 1), containing the complete [5,6,6]-bis-spiroacetal DEF-ring system of the northern hemisphere. (2) (a) Le, L. H.; Erlichman, C.; Pillon, L.; Thiessen, J. J.; Day, A.; Wainman, N.; Eisenhauer, E. A.; Moore, M. J. InVest. New Drugs 2004, 22, 159. (b) Honkanen, R. E.; Golden, T. Curr. Med. Chem. 2002, 9, 2055. (3) For other synthetic studies, see: (a) Liu, J.; Hsung, R. P. Org. Lett. 2005, 7, 2273. (b) Paterson, I.; Anderson, E. A.; Dalby, S. M.; Loiseleur, O.; Abstracts of Papers, 229th National Meeting of the American Chemical Society, San Diego, Mar 13-17, 2005; American Chemical Society: Washington, DC, 2005; ORGN-331. (c) Wang, C.; Forsyth, C. J. Abstracts of Papers, 229th National Meeting of the American Chemical Society, San Diego, Mar 13-17, 2005; American Chemical Society: Washington, DC, 2005; ORGN-414. 4122

Our proposed retrosynthesis of spirastrellolide is outlined in Scheme 1. Simplification of the diastereomer problem might be achieved through a late-stage attachment of the C46 hydroxyl-containing side chain, as in 4, after macrolactonization of an appropriate seco-acid precursor. The DEF-bisspiroacetal subunit may then be isolated as a single region of known relative stereochemistry through disconnection across the C25-C26 bond. This bis-spiroacetal 3 was envisaged to arise from a thermodynamically controlled spiroacetalization of the open-chain precursor 6, following acetonide deprotection. This ketone 6 could itself be derived from the Horner-Wadsworth-Emmons olefination of the aldehyde 7 with phosphonate 8, followed by olefin reduction. Aldehyde 7 may, in turn, be constructed by acetalization of the ketone 9, while phosphonate 8 could arise from a dihydroxylation of the allylic chloride 10, followed by the installation of the dimethyl phosphonate moiety. The synthesis of the C33-C40 aldehyde 7 commenced with the cross-metathesis of buten-3-ol benzyl ether 11 with methyl vinyl acetate 12, using the Grubbs second-generation catalyst (Scheme 2).4 The β,γ-unsaturated ester 13 was obtained in 82% yield as a 4:1 mixture of (E)- and (Z)-isomers, which were further submitted to a Sharpless asymmetric dihydroxylation,5a with concurrent hydroxyldifferentiating lactonization.5b A subsequent recrystallization efficiently removed the minor dihydroxylation byproduct, originating from the (Z)-isomer of 13, while increasing the enantiomeric purity of the desired lactone 145b from 91 to >99% ee. TES protection of alcohol 14, followed by debenzylation and Dess-Martin oxidation, delivered the (4) Chatterjee, A. K.; Choi, T.-L.; Sanders, D. P.; Grubbs, R. H. J. Am. Chem. Soc. 2003, 125, 11360. For an example of cross-metathesis using ester 28, see: Vasbinder, M. M.; Miller, S. J. J. Org. Chem. 2002, 67, 6240. (5) (a) Kolb, H. C.; VanNieuwenhze, M. S.; Sharpless, K. B. Chem. ReV. 1994, 94, 2483. (b) The enantiomer of 14 has previously been reported; see: Garcia, C.; Martin, T.; Martin, V. S. J. Org. Chem. 2001, 66, 1420.

Org. Lett., Vol. 7, No. 19, 2005

aldehyde 15 in 84% yield (three steps), in readiness for a Brown asymmetric crotylation.6

Scheme 2

ity was reduced to 2:1 when a Sharpless asymmetric dihydroxylation was applied using the standard DHQ2PHAL ligand.5a However, by switching the ligand to DHQ2PYR,5a a matched situation occurred and a high selectivity of 97:3 dr was achieved, providing diol 21 in 95% yield.

Scheme 3

In the event, treatment of 15 with (-)-B-(E)-crotyldiisopinocampheylborane provided alcohol 16 in 62% yield, installing the C34 methyl-bearing stereocenter (dr ) 95:5). Oxidation of the alcohol 16 with Dess-Martin periodinane set the stage for the first acetalization, which was performed with a catalytic amount of PPTS in MeOH at 50 °C, providing the methyl acetals 17. Finally, the aldehyde 7 was revealed by ozonolysis of the terminal olefin in 17 (88%, three steps).7 Installation of the challenging C28 chlorine substituent of spirastrellolide employed the asymmetric chloroallylation reaction developed by Oehlschlager (Scheme 3).8 Thus, treatment of the aldehyde 18 with (-)-B-(Z)-chloroallyldiisopinocampheylborane in Et2O at -95 °C provided the allylic chloride 19 in 85% yield with excellent C28/C29 syn stereocontrol (dr > 95:5) and enantiomeric excess (ee > 95%). The C29 hydroxyl was subsequently methylated with Meerwein’s salt, in the presence of proton sponge, to give 20 in 84% yield. With the methyl ether 20 in hand, dihydroxylation of the terminal double bond was investigated. Interestingly, the C28 stereocenter bearing the electronegative chlorine substituent was found to direct dihydroxylation onto the Re-face of the double bond.9 Thus, exposure of 20 to standard Upjohn conditions (OsO4, NMO)10 provided a 6:1 mixture of diols, with the major diastereomer 21 possessing the desired 27,28-anti relationship. Surprisingly, this selectiv(6) Brown, H. C.; Bhat, K. S. J. Am. Chem. Soc. 1986, 108, 293. (7) Some limited epimerization at the sensitive C34 methyl-bearing center took place during acetalization (dr ) 92:8) and ozonolyis (dr ) 88:12). (8) Hu, S.; Jayaraman, S.; Oehlschlager, A. C. J. Org. Chem. 1996, 61, 7513.

Org. Lett., Vol. 7, No. 19, 2005

Following successful dihydroxylation, the diol 21 was protected as the corresponding acetonide. Desilylation with TBAF, followed by Swern oxidation, delivered aldehyde 22 in 83% yield. Treatment of 22 with (MeO)2P(O)CH2Li and subsequent oxidation with Dess-Martin periodinane gave the required β-ketophosphonate 8 (43%). Treatment of aldehyde 7 and phosphonate 8 with Ba(OH)2 in wet THF realized an efficient Horner-WadsworthEmmons fragment coupling to provide the corresponding enone in good yield (85%, Scheme 4).11 The reduction of the CdC bond of this enone was then attempted; however, Pd-catalyzed hydrogenation under a variety of conditions led to decomposition. In contrast, conjugate reduction with Stryker’s reagent12 delivered the ketone 6 smoothly (66%, two steps). The use of wet toluene as a solvent accelerated the reaction and reduced the quantity of reagent required to effect reduction. The stage was now set for investigation of the crucial spiroacetalization reaction. In the event, treatment of ketone 6 with the acidic Dowex 50Wx8 resin, in a mixture of MeOH/water at 70 °C, led to acetonide removal and (9) The same sense of π-facial selectivity is induced by a hydroxyl group in the dihydroxylation of allylic alcohols. See: Cha, J. K.; Christ, W. J.; Kishi, Y. Tetrahedron 1984, 40, 2247. For a discussion of the stereoelectronic effect underlying this selectivity, see: Houk, K. N.; Moses, S. R.; Wu, Y.-D.; Rondan, N. G.; Ja¨ger, V.; Schohe, R.; Fronczek, F. R. J. Am. Chem. Soc. 1984, 106, 3880. (10) VanRheenen, V.; Kelly, R. C.; Cha, D. Y. Tetrahedron Lett. 1976, 14, 1973. (11) Paterson, I.; Yeung, K.-S.; Smaill, J. B. Synlett 1993, 774. (12) Mahoney, W. S.; Brestensky, D. M.; Stryker, J. M. J. Am. Chem. Soc. 1988, 110, 291. 4123

Scheme 4

formation of the tetracyclic bis-spiroacetal product 3 in 40% yield. Under these equilibrating conditions, 3 was obtained as a single diastereomer. The choice of the Dowex resin to mediate this reaction was crucial to the success of the cyclization. Indeed, use of CSA/MeOH led to rapid elimination to form the furan 23, with the acetonide remaining intact. That spiroacetal formation had indeed occurred could be readily confirmed by comparison of the 1H and 13C NMR data of 3 with that of the methyl ester derivative 2 of spirastrellolide.1b As shown by the chart of 1H NMR chemical shift differences between 3 and the same region of spirastrellolide (Scheme 4), there was an excellent match, while the coupling constants were also in good agreement.13 The 13 C NMR spectrum exhibited diagnostic acetal resonances at 98.2 (C31) and 111.6 ppm (C35), also in close agreement with those of the natural product (C31 97.6 ppm, C35 108.7 ppm). In addition, the establishment of the desired spiroacetal configurations at C31 and C35 was reinforced by the observation of strong NOE enhancements between H38 and H27, H38 and H36a, and H36b and both H34 and the C34 methyl substituent. In summary, we have completed a highly convergent synthesis of the fully elaborated chlorinated [5,6,6]-bis-

4124

spiroacetal 3, exploiting asymmetric Sharpless dihydroxylation and allylation methodology. Notably, the advanced intermediate 3 provides additional support for the stereochemical assignment of the DEF region of spirastrellolide1b and bears functionality at C26 and C40 that should allow connection with the remaining C1-C25 and C41-C47 subunits. The following paper14 describes a synthesis of the C1-C25 subunit 5 (see Scheme 1), corresponding to the southern hemisphere of spirastrellolide A. Acknowledgment. We thank the EPSRC (EP/C541677/ 1), Syngenta Crop Protection (secondment of O.L), Merck, and Homerton College, Cambridge (Research Fellowship to E.A.A.), for support and Professor Raymond Andersen (University of British Columbia) for helpful discussions. Supporting Information Available: Spectroscopic data for new compounds. This material is available free of charge via the Internet at http://pubs.acs.org. OL051403C (13) See Supporting Information for full NMR comparisons. (14) Paterson, I.; Anderson, E. A.; Dalby, S. M.; Loiseleur, O. Org. Lett. 2005, 7, 4125.

Org. Lett., Vol. 7, No. 19, 2005